1: \documentstyle[preprint,pra,aps]{revtex}
2: \begin{document}
3: \draft
4: \title{\bf Precise calculation of parity nonconservation
5: in cesium and test of the standard model}
6: \author{V.A. Dzuba, V.V. Flambaum, and J.S.M. Ginges}
7: \address{School of Physics, University of New South Wales,
8: Sydney 2052, Australia}
9: \date{\today}
10: \maketitle
11:
12: \tightenlines
13:
14: %************************************************************
15: \begin{abstract}
16: We have calculated the $6s-7s$ parity nonconserving (PNC) E1 transition
17: amplitude, $E_{PNC}$, in cesium.
18: We have used an improved all-order technique in the calculation
19: of the correlations and have included all significant contributions
20: to $E_{PNC}$.
21: Our final value
22: $E_{PNC}=0.904\Big( 1\pm 0.5\%\Big) \times 10^{-11}iea_{B}(-Q_{W}/N)$
23: has half the uncertainty claimed in old calculations
24: used for the interpretation of Cs PNC experiments.
25: The resulting nuclear weak charge $Q_{W}$ for Cs deviates by
26: about $2\sigma$ from the value predicted by the standard model.
27:
28: \end{abstract}
29: \vspace{1cm}
30:
31: \pacs{PACS: 11.30.Er, 12.15.Ji, 32.80.Ys, 31.30.Jv}
32: %************************************************************
33:
34: \section{Introduction}
35:
36: There is an ongoing discussion on whether measurements of parity
37: nonconservation (PNC) in atoms can be used to study new physics beyond the
38: standard model (for a history of PNC in atoms, see, e.g.,
39: the book \cite{khriplovich} or review \cite{bouchiat}).
40: Such a possibility relies on the accuracy of
41: the measurements and the theoretical analysis.
42: The best accuracy so far has been achieved for the cesium atom.
43: In 1997, the Boulder group measured the PNC amplitude in cesium
44: to an accuracy of 0.35\% \cite{wood97}. The best calculations available
45: at that time were published by the Novosibirsk group in 1989 \cite{dzuba89}
46: and the Notre Dame group in 1990 \cite{blundell90}.
47: Both works claimed an accuracy of 1\%.
48: This claim of 1\% accuracy was based in part on the comparison
49: of the experimental and theoretical values relevant to the PNC amplitude
50: (energies, electromagnetic E1 transition amplitudes, hyperfine structure).
51: Later, new, more accurate measurements (see, e.g. \cite{rafac,bennett99})
52: led to better agreement between theory and experiment for the
53: E1 transition amplitudes.
54: This allowed Bennett and Wieman to suggest that the actual
55: accuracy of the PNC calculations in cesium is 0.4\% \cite{wieman99}.
56: With this accuracy, the value of the nuclear weak charge $Q_{W}$
57: of the cesium atom which follows from the measurements \cite{wood97} and
58: calculations \cite{dzuba89,blundell90} deviates from the standard model
59: value by $2.5\sigma$ \cite{wieman99}.
60: The implications of this deviation for physics beyond the standard model
61: have been examined in several works \cite{RC,JR,EL}.
62:
63: This result generated many recent works revisiting calculations of
64: PNC in cesium.
65: Derevianko \cite{derevianko00} demonstrated that the contribution
66: of the Breit interaction to the PNC amplitude $E_{PNC}$ (-0.6\%) is
67: substantially larger than previous estimates and reduces the deviation
68: from the standard model. The Breit contribution was neglected in \cite{dzuba89}
69: and underestimated in \cite{blundell90}. The result of Derevianko was
70: later confirmed in independent calculations \cite{harabati,kozlov01}.
71: A new many-body calculation of $E_{PNC}$ was performed by
72: Kozlov {\it et al.} \cite{kozlov01}.
73: It was calculated in second-order in the residual interaction
74: with averaged screening factors.
75: The result is in excellent agreement with similar calculations by
76: Blundell {\it et al.} \cite{blundell90} and differs by less
77: than 0.5\% from the more complete (``all-orders'') calculations
78: \cite{dzuba89,blundell90}.
79:
80: Radiative corrections to the weak charge
81: of order $\alpha$ were calculated a long time ago
82: in \cite{marciano} (see also \cite{lynnbed}).
83: However, there are important ``strong field'' radiative corrections
84: of order $Z\alpha ^{2}$ and $Z^2\alpha ^{3} \ln^2(\lambda/R_n)$
85: which have recently been considered in
86: \cite{milstein} (note that the latter correction is larger).
87: Here $\lambda$ is the electron Compton wavelength, and
88: $R$ is the nuclear radius.
89: The main contribution to $E_{PNC}$ was found to be about $0.4\%$.
90: This contribution originates from the radiative corrections to the
91: weak matrix element due to the Uehling potential
92: (recently this contribution was calculated numerically in \cite{johnson01}).
93:
94: Among other corrections considered
95: one should mention a correction for the neutron distribution
96: \cite{derevianko} which is quite small: -0.2\% of $E_{PNC}$.
97:
98: Note that Breit, Uehling, and neutron skin corrections are
99: smaller than the uncertainty of 1\% claimed for the calculated $E_{PNC}$
100: amplitude in works \cite{dzuba89,blundell90}. This uncertainty was
101: mostly associated with the accuracy of the calculations of correlation
102: corrections. Therefore, there is a very important question as to
103: whether calculations of correlations were, or can be, performed to
104: better than 1\% accuracy.
105: The analysis of the accuracy involves a comparison of
106: calculated and measured atomic quantities
107: {\em and} some additional ``internal'' tests, such as, e.g.,
108: checking the stability of the results against variation of certain
109: parameters.
110: Comparison with experiment alone is an important
111: but not sufficient part of the analysis. This is especially true when
112: better accuracy is claimed for calculations performed many years
113: ago and all the details can hardly be reconstructed. In our view,
114: the only reliable way to improve the accuracy is to do the calculations
115: again, trying to improve the method and numerical procedures at every
116: stage, and repeat the analysis of the accuracy.
117: That is what we do in the present paper.
118:
119: We have made several important improvements to the
120: method developed in 1989 \cite{dzuba89}. First, we calculated a new series
121: of higher-order correlation diagrams which account for the effect of
122: screening of the Coulomb interaction in the exchange correlation diagram.
123: In our earlier work \cite{dzuba89} the screening effect was calculated
124: for the direct correlation diagram only, while screening factors
125: were used for the exchange diagram.
126: The values of these factors were found by looking at the effect of
127: screening in the direct diagram; however, the effect
128: of screening in the exchange diagram may differ.
129: In our present work we calculate the effect of screening for both diagrams.
130:
131: Second, we have made significant improvements to the numerical procedures
132: at every stage of the calculations. This involves, for instance,
133: using a more dense coordinate grid, including more terms
134: into the summation over virtual states,
135: and restoring the lower Dirac component of the wave function everywhere.
136: We have also used a more accurate nuclear charge distribution.
137:
138: Our result for the PNC amplitude in cesium
139: (without Breit, radiative, and neutron distribution corrections)
140: $E_{PNC}=0.908\Big( 1\pm 0.5\% \Big) \times 10^{-11}iea_{B}(-Q_{W}/N)$
141: coincides exactly with the old result
142: \cite{dzuba89} but has a much smaller uncertainty.
143: ($N$ is the neutron number.)
144: To avoid misunderstanding, we should note
145: that particular contributions to this value are slightly different
146: in the old and the new calculations (for example, the Hartree-Fock
147: value of the amplitude has changed by 0.3\%).
148: Therefore, we are indeed talking about a new result.
149:
150: We have also calculated the contributions of the
151: Breit interaction (-0.61\%) and the Uehling potential (0.40\%)
152: to $E_{PNC}$. There is very good agreement between different calculations
153: for these values and they can be considered well-established.
154: The resulting value for the PNC amplitude is
155: $E_{PNC}=0.906\Big( 1\pm 0.5\% \Big)\times 10^{-11}iea_{B}(-Q_{W}/N)$.
156: The neutron-distribution correction,
157: estimated as $(-0.2 \pm 0.1)\%$, reduces our final value to
158: $E_{PNC}=0.904\Big( 1\pm 0.5\% \Big) \times 10^{-11}iea_{B}(-Q_{W}/N)$.
159:
160: %************************************************************
161: \section{Method of calculation}
162: \label{sec:method}
163:
164: The method we have used here is very similar to that used in our 1989
165: work \cite{dzuba89}.
166: As a zero approximation we use the relativistic Hartree-Fock
167: method. The perturbation in our approach is the residual
168: interaction - the difference between the exact and Hartree-Fock
169: Hamiltonians.
170: For summing of diagrams we use a combination of the ``correlation potential
171: method'' \cite{dzuba87}, which is a way to treat correlations by introducing
172: a single-electron operator (correlation potential) $\hat \Sigma$,
173: and ``perturbation theory in the screened electron residual interaction''
174: \cite{dzuba89energy,dzuba89e1hfs}, which is used to calculate $\hat \Sigma$.
175: The method is an all-order technique, in terms of treating correlations,
176: and is not a version of the popular coupled-cluster approach.
177: The dominating sequences of higher-order correlation diagrams
178: correspond to real physical phenomena like collective screening of the
179: Coulomb interaction between electrons and the hole-particle interaction.
180: They are included in all orders in our technique.
181: An important strong point of the method is that its complexity does not go
182: beyond the calculation of the correlation potential $\hat \Sigma$.
183: When $\hat \Sigma$ is ready, the calculation of energies and
184: matrix elements is relatively simple.
185: $\hat{\Sigma}$ is used to calculate single-electron Brueckner orbitals.
186: The calculation of matrix elements with Brueckner orbitals
187: already includes most of the correlations.
188: There are additional ``non-Brueckner'' contributions to matrix elements such
189: as structural radiation and normalization of many-electron states. These
190: contributions are small for $s$ and $p$ states and in most cases they
191: can be expressed in terms of derivatives of $\hat \Sigma$.
192: This makes the calculation of hyperfine structure, transition amplitudes,
193: etc. similar to the calculation of energies.
194:
195: The method works very well for alkaline atoms. The calculated
196: energy levels and transition amplitudes deviate by a fraction
197: of a percent from experiment.
198: The hyperfine structure (hfs) of $s$ and $p$ states has an
199: accuracy of better than 1\%.
200:
201: \subsection{Correlation potential method}
202:
203: The Cs atom has one electron above closed shells. Therefore, it is natural
204: to start the calculations from the relativistic Hartree-Fock (RHF) method
205: in the $\hat{V}^{N-1}$ approximation (one electron in the field of the
206: $N-1$ core electrons).
207: The single-electron RHF Hamiltonian is
208: \begin{equation}
209: \label{eq:RHF}
210: \hat{H}_{0}=c\mbox{\boldmath$\alpha$}\cdot \hat{{\bf p}}+(\beta -1)c^{2}-
211: Z\alpha/r+\hat{V}^{N-1} \ ,
212: \end{equation}
213: $\mbox{\boldmath$\alpha$}$ and $\beta$ are Dirac matrices and
214: $\hat{{\bf p}}$ is the electron momentum.
215: For cesium the accuracy of the RHF energies is of the order of $10\%$.
216:
217: To improve the accuracy one needs to include correlations.
218: In order to do so we introduce a ``correlation potential'' $\hat \Sigma$
219: \cite{dzuba87}. $\hat \Sigma$ is a single-electron energy-dependent
220: non-local operator defined in such a way that its average value over
221: a single-electron state $a$ is the correlation correction to the energy
222: of this state,
223: \begin{equation}
224: \label{eq:S}
225: \Delta \epsilon_a = \langle a | \hat \Sigma(\epsilon_a) | a \rangle.
226: \end{equation}
227: The calculation of $\hat \Sigma$ will be described in the next section.
228: $\hat \Sigma$ is added to the Hartree-Fock Hamiltonian to calculate
229: single-electron states of the external electron,
230: \begin{equation}
231: \label{eq:Br}
232: (\hat H_0 + \hat \Sigma - \epsilon) \psi^{\rm Br} = 0.
233: \end{equation}
234: There are two major benefits from inclusion of $\hat \Sigma$ in the equations.
235: First, the iteration of $\hat \Sigma$ is an important higher-order effect
236: which gives a significant contribution to the energy and wave function.
237: Second, by solving Eq.~(\ref{eq:Br}) we obtain single-electron orbitals
238: which are called Brueckner orbitals which already include correlations.
239: These orbitals can then be used to calculate matrix elements for
240: hfs, PNC, etc.
241:
242: When appropriate higher-order diagrams are included into the calculation
243: of $\hat \Sigma$, as will be described in the next section, by solving
244: Eq.~(\ref{eq:Br}) we obtain energies of $s$ and $p$ states of alkaline
245: atoms with an accuracy of about 0.1\%. This is a radical improvement
246: over the 10\% accuracy of the Hartree-Fock approximation.
247:
248: To accurately calculate the interaction of external fields with atomic
249: electrons one needs also to take into account the core polarization effect.
250: We do this using the time-dependent Hartree-Fock (TDHF) method
251: (which is equivalent to the random-phase approximation with exchange).
252: In this method every single-electron orbital has the form
253: \[ \psi_n + \delta \psi_n \ , \] where $\psi_n$ is the unperturbed
254: Hartree-Fock orbital (or Brueckner one for states above the core)
255: and $\delta \psi_n$ is a correction to the orbital caused by
256: the external field.
257: To find $\delta \psi_n$ one needs to solve the equations
258: \begin{equation}
259: \label{eq:RPA}
260: (\hat H_0 - \epsilon)\delta \psi_n = - (\delta \epsilon +
261: \hat h + \delta \hat V_h) \psi_n.
262: \end{equation}
263: Here $\hat h$ is the operator of the external field, $\delta \hat V_h$
264: is the correction to the Hartree-Fock potential due to modification
265: of the core states in the external field, and
266: $\delta \epsilon$ is the correction to the energy.
267:
268: Equations (\ref{eq:RPA}) are first solved self-consistently for core
269: states to obtain $\delta \hat V_h$. Then Brueckner orbitals are used
270: to calculate matrix elements between states of the external electron
271: \begin{equation}
272: \label{eq:me}
273: M_{ab} = \langle \psi^{\rm Br}_b |\hat{h} +\delta \hat{V}_{h}|
274: \psi^{\rm Br}_a \rangle.
275: \end{equation}
276: This expression includes most of the correlations and already gives hfs
277: and transition amplitudes in alkaline atoms to an accuracy close to 1\%.
278: There are also contributions arising from the normalization of states and
279: structural radiation \cite{dzuba87}. When these contributions are included
280: the accuracy improves.
281:
282: \subsection{Calculation of the correlation potential}
283:
284: We use the Feynman diagram technique to calculate the correlation potential
285: $\hat \Sigma$. The main reason for this is that this technique is more
286: convenient for inclusion of higher-order diagrams. The drawback of it
287: is the necessity to integrate over frequencies numerically.
288:
289: In the lowest (second) order in the residual Coulomb interaction,
290: $\hat \Sigma$ is given by the direct and exchange diagrams in
291: Fig.~\ref{fig:Sigma2}.
292: The corresponding mathematical expressions are
293: \begin{eqnarray}
294: \hat \Sigma^{(2)}_{\rm d} (\epsilon, r_i, r_j) &=&
295: \sum_{nm} \int \frac{d\omega}{2\pi} G_{ij}(\epsilon+\omega)
296: \hat Q_{im}\hat \Pi_{mn}(\omega)\hat Q_{nj} \label{eq:sigma2d},\\
297: \hat \Sigma^{(2)}_{\rm ex} (\epsilon, r_i, r_j) &=&
298: \sum_{mn} \int \int \frac{d\omega_1}{2\pi}\frac{d\omega_2}{2\pi}
299: \hat Q_{in}G_{im}(\epsilon+\omega_1)G_{mn}(\epsilon+\omega_1+\omega_2)
300: \hat Q_{mj}G_{nj}(\epsilon+\omega_2) \label{eq:sigma2ex},
301: \end{eqnarray}
302: where summation over $m$ and $n$ is a numerical implementation of the
303: integration over $r_m$ and $r_n$, $\hat Q$ is the Coulomb interaction,
304: $G$ is the Hartree-Fock Green function, $\hat \Pi$ is the
305: polarization operator:
306: \begin{eqnarray}
307: &\hat Q_{ij} = \frac{e^2}{|r_i - r_j|}, \label{eq:q}\\
308: &G_{ij}(\epsilon) =
309: \sum_{\gamma} \frac{|\gamma\rangle_i\langle\gamma|_j}
310: {\epsilon-\epsilon_{\gamma}+i\delta} +
311: \sum_{n} \frac{|n\rangle_i\langle n|_j}
312: {\epsilon-\epsilon_{n}-i\delta}, \ \ \delta \rightarrow 0,
313: \label{eq:gf}\\
314: &\hat \Pi_{ij}(\omega)=\int \frac{d\alpha}{2\pi}G_{ij}(\omega+\alpha)
315: G_{ij}(\alpha) =\sum_n\psi_n^{\dagger}\big(
316: G_{ij}(\epsilon_n+\omega)+G_{ij}(\epsilon_n-\omega)\big)\psi_n.
317: \label{eq:po}
318: \end{eqnarray}
319: Here $|n \rangle$ are core RHF states and $|\gamma \rangle$ are RHF excited
320: states;
321: $\delta$ shows the pole passing rule for integration over frequencies.
322: (Let us remind the reader that we use the Feynman diagram
323: technique to treat a basically non-relativistic correlation problem, e.g.,
324: our polarization operator (\ref{eq:po}) represents the polarization
325: of the atomic core, not the vacuum.)
326:
327: The integration over frequencies in expressions (\ref{eq:sigma2d}) and
328: (\ref{eq:sigma2ex})
329: can easily be performed analytically, reducing the diagrams in
330: Fig.~\ref{fig:Sigma2}
331: to the four usual Goldstone diagrams presented, e.g.,
332: in work \cite{dzuba87}.
333: When higher-orders are included in $\hat \Sigma$,
334: only the integration for the polarization operator (\ref{eq:po}) can still be
335: done analytically. The other integrals have to be calculated numerically.
336:
337: As was demonstrated in our earlier work \cite{dzuba89energy},
338: the most important higher-order correlation contributions to
339: $\hat \Sigma$ come from screening
340: of the Coulomb interaction between atomic electrons by other electrons
341: and from the hole-particle interaction in the polarization operator (another
342: important higher-order effect is the iteration of $\hat \Sigma$, which is
343: included separately, by inserting $\hat \Sigma$ into equations for
344: single-electron orbitals; see previous section).
345:
346: The hole-particle interaction in the polarization operator $\hat \Pi$
347: (Fig.~\ref{fig:h-p}) is included by replacing the $\hat{V}^{N-1}$
348: Hartree-Fock potential
349: in the calculation of $\hat \Pi$ by the $\hat{V}^{N-2}$ potential. Since
350: the calculation involves iterations of the RHF equation, the hole-particle
351: interaction is included in all orders.
352:
353: The screening of the Coulomb interaction is a more complicated effect.
354: The corresponding higher-order diagrams can be obtained from the second-order
355: diagrams (Fig.~\ref{fig:Sigma2}) by the insertion of a series of hole-particle
356: loops into every Coulomb line. Therefore, instead of,
357: e.g., one direct diagram (Fig.~\ref{fig:Sigma2}) we have an infinite chain of
358: higher-order diagrams (Fig.~\ref{fig:Sigmad}). One can see that an internal
359: part of this chain (before the integration over $\omega$) forms a matrix
360: geometric progression
361: \[
362: \hat \Pi + \hat \Pi \hat Q \hat \Pi +
363: \hat \Pi \hat Q \hat \Pi \hat Q \hat \Pi + \dots
364: \]
365: The sum of this progression is
366: \begin{equation}
367: \hat \pi(\omega)=
368: \hat \Pi(\omega)\big[ 1-\hat Q \hat \Pi(\omega)\big]^{-1},
369: \label{eq:screenedpi}
370: \end{equation}
371: and can be called the ``screened polarization operator''. Screening
372: of the Coulomb interaction in the direct correlation diagram is included by
373: replacing the unscreened polarization operator $\hat \Pi$ in
374: Eq.~(\ref{eq:sigma2d}) by the screened polarization operator $\hat \pi$:
375: \begin{eqnarray}
376: \label{eq:sigmad}
377: \hat \Sigma_{\rm d} (\epsilon, r_i, r_j) &=&
378: \sum_{nm} \int \frac{d\omega}{2\pi} G_{ij}(\epsilon+\omega)
379: \hat Q_{im}\hat{\pi}_{mn}(\omega)\hat Q_{nj}.
380: \end{eqnarray}
381:
382: It is convenient also to introduce an operator of the screened
383: Coulomb interaction
384: \begin{equation}
385: \widetilde Q(\omega)=\hat Q (1-\hat Q \hat{\pi}(\omega))^{-1}.
386: \label{eq:screenedQ}
387: \end{equation}
388: This expression can be obtained in a similar way to the screened
389: polarization operator $\hat \pi$ by summing a matrix geometric progression
390: corresponding to the chain of diagrams in Fig.~\ref{fig:screeninghp}.
391: An expression for the direct correlation diagram can be re-written in
392: terms of $\widetilde Q$:
393: \begin{eqnarray}
394: \label{eq:sigmade}
395: \hat \Sigma_{\rm d} (\epsilon, r_i, r_j) &=&
396: \sum_{nm} \int \frac{d\omega}{2\pi} G_{ij}(\epsilon+\omega)
397: \hat Q_{im}\hat{\Pi}_{mn}(\omega)\widetilde Q_{nj}(\omega).
398: \end{eqnarray}
399: The screened Coulomb interaction operator $\widetilde Q$ can also
400: be used to include screening into the exchange correlation diagram.
401: This is done by substituting $\hat Q \rightarrow \widetilde Q(\omega)$ in
402: equation~(\ref{eq:sigma2ex})
403: \begin{eqnarray}
404: \label{eq:sigmaex}
405: &&\hat \Sigma_{\rm ex} (\epsilon, r_i, r_j) = \nonumber \\
406: &&\sum_{mn} \int \int \frac{d\omega_1}{2\pi}\frac{d\omega_2}{2\pi}
407: \widetilde Q_{in}(\omega_1)G_{im}(\epsilon+\omega_1)
408: G_{mn}(\epsilon+\omega_1+\omega_2)
409: \widetilde Q_{mj}(\omega_2)G_{nj}(\epsilon+\omega_2).
410: \end{eqnarray}
411:
412: The second-order correlation potential with the
413: hole-particle interaction and screening of the Coulomb interaction
414: included in all orders is depicted in Fig.~\ref{fig:Sigmahpsc}.
415:
416: In our 1989 work \cite{dzuba89} we used an approximate expression for the
417: screened Coulomb interaction in the exchange correlation diagram
418: \begin{equation}
419: \label{eq:fk}
420: \widetilde Q_k \approx f_k \hat Q_k,
421: \end{equation}
422: where $k$ is the multipolarity of the Coulomb interaction and $f_k$ is the
423: screening factor. In this expression the dependence of the screening on
424: frequency is neglected. The values of the screening factors were obtained by
425: calculating the direct diagram with and without screening.
426: The use of average screening factors which do not depend on frequency
427: significantly simplified the calculation of the exchange diagram. Like
428: in pure second-order, the Goldstone diagram technique and direct summation
429: over a complete set of single-electron states were used. No integration
430: over frequencies was needed.
431:
432: In the present work we do the full-scale calculation of the exchange diagram
433: using Eq.~(\ref{eq:sigmaex}). This makes the method totally
434: {\it ab initio} since no screening factors are used. The drawback of
435: the method is the need to do double integration over frequencies numerically
436: (note that there is only a single numerical integration over frequencies
437: in the direct diagram).
438:
439: \subsection{PNC amplitude}
440:
441: To calculate the PNC $6s-7s$ amplitude in Cs one needs to include
442: two external fields acting on the atomic electrons:
443: the weak field of the nucleus and the electric dipole
444: (E1) field of the photon.
445: The nuclear spin-independent weak interaction of an electron with
446: the nucleus is
447: \begin{equation}
448: \label{eq:h_w}
449: \hat{H}_{W}=\frac{G_{F}}{2\sqrt{2}}\rho (r)Q_{W}\gamma _{5}
450: \end{equation}
451: where $G_{F}$ is the Fermi constant, $Q_{W}$ is the weak charge of the
452: nucleus, $\gamma _{5}$ is a Dirac matrix, and $\rho (r)$ is the nuclear
453: density. The E1 Hamiltonian is
454: \begin{equation}
455: \label{eq:E1}
456: \hat H_{E1} = - {\bf d \cdot E}(e^{-i\omega t}+e^{i\omega t}),
457: \end{equation}
458: where ${\bf d}$ is the dipole moment operator.
459:
460: In the TDHF method, a single-electron wave function is
461: \begin{equation}
462: \label{eq:wf}
463: \psi = \psi_0 + \delta \psi + Xe^{-i\omega t}+Ye^{i\omega t}+
464: \delta Xe^{-i\omega t}+ \delta Ye^{i\omega t},
465: \end{equation}
466: where $\delta \psi$ is the correction due to the weak interaction
467: acting alone, $X$ and $Y$ are corrections due to the photon field
468: acting alone, and $\delta X$ and $\delta Y$ are corrections due
469: to both fields acting simultaneously.
470: These corrections are found by solving self-consistently the system
471: of the TDHF equations for the core states
472: \begin{eqnarray}
473: (\hat{H}_{0}-\epsilon)\delta \psi &=&
474: -(\hat{H}_{W}+\delta \hat{V}_{W})\psi, \label{eq:WE1:1}\\
475: (\hat{H}_{0}-\epsilon -\omega)X &=&
476: -(\hat{H}_{E1}+\delta \hat{V}_{E1})\psi, \label{eq:WE1:2}\\
477: (\hat{H}_{0}-\epsilon +\omega)Y&=&
478: -(\hat{H}_{E1}^{\dagger}+\delta \hat{V}_{E1}^{\dagger})\psi,
479: \label{eq:WE1:3}\\
480: (\hat{H}_{0}-\epsilon -\omega)\delta X &=&
481: -\delta \hat{V}_{E1}\delta \psi
482: -\delta \hat{V}_{W}X
483: -\delta \hat{V}_{E1W}\psi, \label{eq:WE1:4}\\
484: (\hat{H}_{0}-\epsilon +\omega)\delta Y&=&
485: -\delta \hat{V}_{E1}^{\dagger}\delta \psi
486: -\delta \hat{V}_{W}^{\dagger}Y
487: -\delta \hat{V}_{E1W}^{\dagger} \psi,\label{eq:WE1:5}
488: \end{eqnarray}
489: where $\delta \hat{V}_W$ and $\delta \hat{V}_{E1}$ are corrections to the
490: core potential due to the weak and E1 interactions, respectively,
491: and $\delta \hat{V}_{E1W}$ is the correction to the core potential
492: due to the simultaneous action of the weak field and the electric field
493: of the photon.
494:
495: The TDHF contribution to $E_{PNC}$ between the states $6s$ and $7s$
496: is given by
497: \begin{equation}
498: \label{eq:TDHF}
499: E_{PNC}^{TDHF}=\langle \psi _{7s}|\hat{H}_{E1}+\delta \hat{V}_{E1}|
500: \delta \psi _{6s}\rangle +
501: \langle \psi _{7s}|\hat{H}_{W}+\delta \hat{V}_{W}|
502: X _{6s}\rangle +
503: \langle \psi _{7s}|\delta \hat{V}_{E1W}|\psi _{6s}\rangle \ .
504: \end{equation}
505: The corrections $\delta \psi_{6s}$ and $X_{6s}$ are found by solving
506: the equations~(\ref{eq:WE1:1}-\ref{eq:WE1:2}) in the field of the
507: frozen core (of course, the amplitude (\ref{eq:TDHF}) can instead
508: be expressed in terms of corrections to $\psi _{7s}$).
509:
510: If we use Brueckner orbitals instead of RHF orbitals to calculate
511: the PNC amplitude in Eq.~(\ref{eq:TDHF}) we include all-orders
512: in $\hat{\Sigma}$ contributions to $E_{PNC}$.
513: However, the correlation potential is energy-dependent,
514: $\hat{\Sigma}=\hat{\Sigma}(\epsilon)$, which means that $\hat \Sigma$
515: operators for the $6s$ and $7s$ states are different.
516: We should consider the proper energy-dependence at least in first-order
517: in $\hat \Sigma$ (higher-order corrections are small and the proper
518: energy-dependence is not important for them).
519: The first-order in $\hat{\Sigma}$ corrections to $E_{PNC}$ are
520: presented diagrammatically in Fig.~\ref{fig:pncdom}.
521: We can write these as
522: \begin{equation}
523: \label{eq:pnc-cor}
524: \langle \psi _{7s}|\hat{\Sigma}_s(\epsilon_{7s})|\delta X_{6s}\rangle
525: +\langle \delta\psi _{7s}|\hat{\Sigma}_p(\epsilon_{7s})|X_{6s}\rangle
526: +\langle \delta Y_{7s}|\hat{\Sigma}_s(\epsilon_{6s})|\psi_{6s}\rangle
527: +\langle Y_{7s}|\hat{\Sigma}_p(\epsilon_{6s})|\delta \psi_{6s}\rangle \ .
528: \end{equation}
529: The non-linear in $\hat{\Sigma}$ contribution can be found by subtracting
530: from the all-orders result the first-order value found in the same method.
531:
532: The correlation corrections to $E_{PNC}$ we have considered so far
533: are usually called ``Brueckner-type'' corrections.
534: (In this case the external field interacts with the external electron
535: lines.) There are also contributions to $E_{PNC}$ in which the
536: external field acts inside the correlation potential
537: (see Fig.~\ref{fig:pncint}).
538: Those diagrams in which the E1 interaction occurs in the internal
539: lines are known as ``structural radiation'',
540: while those in which the weak interaction occurs in the internal lines are
541: known as the ``weak correlation potential''.
542: There is another second-order correction to the amplitudes which arises
543: from the normalization of states \cite{dzuba87}.
544: The structural radiation, weak correlation potential, and
545: normalization contributions are suppressed by the small parameter
546: $E_{\rm ext}/E_{\rm core}\sim 1/10$, where $E_{\rm ext}$ and
547: $E_{\rm int}$ are excitation energies of the external and core electrons,
548: respectively.
549:
550: We have also included a correction to $E_{PNC}$ due to the Breit interaction.
551: We calculated this in a way similar to the earlier work \cite{harabati}.
552: The main difference is that in the present calculations we have also
553: included the Breit contribution to the last term in Eq.~(\ref{eq:TDHF}).
554: This makes the calculations more consistent but doesn't change the result
555: significantly.
556:
557: We postpone the analysis of radiative corrections until
558: Section~\ref{sec:rad}.
559:
560: %************************************************************
561: \section{$6s-7s$ PNC amplitude}
562: \label{sec:results}
563:
564: The results of our calculation for the $6s-7s$ PNC amplitude
565: are presented in Table~\ref{tab:pnci}.
566: Notice that the time-dependent Hartree-Fock value gives a
567: contribution to the total amplitude of about $98\%$.
568: The point is that there is a strong cancellation of the
569: correlation corrections to the PNC amplitude.
570: The stability of the PNC amplitude compared to other
571: quantities in which the correlation corrections are large
572: will be discussed in more detail in Section~\ref{sec:accuracy}.
573: Notice that the values do not differ significantly from our 1989 results
574: (compare the 1989 final result $0.908\times 10^{-11}iea_{B}(-Q_{W}/N)$
575: with ``Subtotal'' of Table~\ref{tab:pnci} for the current calculation).
576: The new series of higher-order diagrams and higher numerical accuracy
577: of the current work has therefore not changed the previous result
578: (note, however, that particular contributions are slightly different).
579:
580: The mixed-states approach has also been performed in
581: \cite{blundell90} and \cite{kozlov01} to determine the PNC amplitude
582: in cesium.
583: However, in these works the screening of the
584: electron-electron interaction was included in a simplified way.
585: In \cite{blundell90} empirical screening factors were placed before
586: the second-order correlation corrections $\hat{\Sigma}^{(2)}$ to fit the
587: experimental values of energies.
588: Kozlov {\it et al.} \cite{kozlov01} introduced screening factors
589: based on average screening factors calculated for the Coulomb
590: integrals between valence electron states.
591: The results obtained by these groups
592: (without the Breit interaction, i.e., corresponding to the
593: Subtotal of Table~\ref{tab:pnci})
594: are $0.904$ \cite{blundell90} and $0.905$ \cite{kozlov01}.
595: To be sure that we understand the difference between these values and
596: our value, we performed a pure second-order
597: (i.e., using $\hat{\Sigma}^{(2)}$) calculation and fitted the energies
598: (as was done in \cite{blundell90}) and reproduced their result, $0.904$.
599:
600: In the work \cite{blundell90} a calculation using the
601: sum-over-states method was also performed.
602: In the sum-over-states approach the $6s-7s$ PNC amplitude
603: is expressed in the form
604: \begin{equation} \label{sum}
605: E_{PNC}=\sum _{n} \Big(
606: \frac{\langle 7s |\hat{H}_{W}|np\rangle \langle np|\hat{H}_{E1}|6s \rangle}
607: {E_{7s}-E_{np}} +
608: \frac{\langle 7s |\hat{H}_{E1}|np\rangle \langle np|\hat{H}_{W}|6s \rangle}
609: {E_{6s}-E_{np}} \Big) \ .
610: \end{equation}
611: The authors of reference \cite{blundell90} include single, double, and
612: selected triple excitations into their wave functions.
613: Note, however, that even if wave functions of $6s$, $7s$, and intermediate
614: $np$ states are calculated exactly
615: (i.e., with all configuration mixing included)
616: there are still some missed contributions in this approach.
617: Consider, e.g., the intermediate state $6p\equiv 5p^{6}6p$.
618: It contains an admixture of states $5p^{5}ns6d$:
619: $\widetilde{6p}=5p^{6}6p + \alpha 5p^{5}ns6d+...$.
620: This mixed state is included into the sum (\ref{sum}).
621: However, the sum (\ref{sum}) must include all many-body states
622: of opposite parity.
623: This means that the state $\widetilde{5p^{5}ns6d}=
624: 5p^{5}ns6d- \alpha 5p^{6}6p+...$ should also be included into
625: the sum. Such contributions to $E_{PNC}$ have never
626: been estimated directly within the sum-over-states approach.
627: However, they are included into our mixed-states calculation.
628: The result of the sum-over-states approach, 0.909,
629: is very close to the result of the mixed-states approach, 0.908.
630: It is important to note that the omitted higher-order many-body corrections
631: are different in these two methods.
632: This may be considered as an argument that the omitted many-body corrections
633: in both calculations are small.
634: Of course, here we assume that the omitted many-body corrections to both
635: values (which, in principle, are completely different) do not
636: ``conspire'' to give exactly the same magnitude.
637:
638: Therefore we will take $0.908$ for the value of $E_{PNC}$
639: (Subtotal of Table~\ref{tab:pnci}) as this corresponds
640: to the most complete mixed-states calculation
641: and is in agreement with the sum-over-states calculation of
642: reference \cite{blundell90}.
643:
644: With Breit, our result becomes $0.902\times 10^{-11}iea_{B}(-Q_{W}/N)$.
645: This correction is in agreement with
646: \cite{derevianko00,harabati,kozlov01}.
647:
648: We use the two-parameter Fermi model for the proton and neutron distributions:
649: \begin{equation}
650: \rho (r)=\rho _{0}\Big[ 1+\exp [(r-c)/a] \Big] ^{-1}\ ,
651: \end{equation}
652: where $t=a(4\ln 3)$ is the skin-thickness,
653: $c$ is the half-density radius, and
654: $\rho _{0}$ is found from the normalization condition $\int \rho dV=1$.
655: In 1989 the thickness and half-density radius for the proton
656: distribution were taken to be $t_{p}=2.5~{\rm fm}$ and
657: $c_{p}=5.6149~{\rm fm}$ (corresponding to a root-mean-square
658: (rms) radius $\langle r_{p}^{2}\rangle ^{1/2}=4.836~{\rm fm}$).
659: In this work we have used improved parameters
660: $t_{p}=2.3~{\rm fm}$ and $c_{p}=5.6710~{\rm fm}$
661: ($\langle r_{p}^{2}\rangle ^{1/2}=4.804~{\rm fm}$) \cite{fricke95}.
662: This changes the wave functions slightly,
663: leading to a very small correction to the PNC amplitude
664: of $0.08\%$ ($0.0007$).
665: (This is in agreement with a simple analytical estimate:
666: the factor accounting for the change in the electron density
667: is $\sim (4.804/4.836)^{-Z^{2}\alpha ^{2}}\sim 0.1\% \ $.)
668: This correction has already been included into the TDHF value.
669:
670: In the work \cite{dzuba89} we used the proton distribution in the
671: weak interaction Hamiltonian (Eq.~(\ref{eq:h_w})).
672: In the current work we have found the small correction to
673: $E_{PNC}$ which arises from taking the (poorly understood)
674: neutron density in Eq.~(\ref{eq:h_w}).
675: We use the result of Ref. \cite{r_{np}} for the difference
676: $\Delta r_{np}=0.13(4)~{\rm fm}$
677: in the root-mean-square radii of the neutrons
678: $\langle r_{n}^{2}\rangle ^{1/2}$ and protons
679: $\langle r_{p}^{2}\rangle ^{1/2}$.
680: We have considered three cases which correspond to the same value of
681: $\langle r_{n}^{2}\rangle$: (i) $c_{n}=c_{p}$, $a_{n}>a_{p}$;
682: (ii) $c_{n}>c_{p}$, $a_{n}>a_{p}$; and (iii) $c_{n}>c_{p}$, $a_{n}=a_{p}$
683: (using the relation $\langle r_{n}^{2}\rangle \approx \frac{3}{5}c_{n}^{2}
684: +\frac{7}{5}\pi ^{2}a_{n}^{2}$).
685: We have found that $E_{PNC}$ shifts from $-0.18\%$ to $-0.21\%$
686: when moving from the extreme $c_{n}=c_{p}$ to the extreme $a_{n}=a_{p}$.
687: Therefore, $E_{PNC}$ changes by about $-0.2\%$ ($-0.0018$)
688: due to consideration of the neutron distribution.
689: This is in agreement with Derevianko's estimate,
690: $-0.19(8)\%$ \cite{derevianko}.
691:
692: In the next section we discuss the radiative corrections to
693: $E_{PNC}$. The dominating contribution comes from the Uehling potential and
694: increases the amplitude by 0.4\%.
695:
696: Therefore, we have
697: \begin{equation}
698: E_{PNC}=0.9041 \times 10^{-11}iea_{B}(-Q_{W}/N)
699: \end{equation}
700: as our central point for the PNC amplitude.
701: The error will be estimated in Section~\ref{sec:accuracy}.
702:
703: %*****************************************************************
704: \section{QED-type radiative corrections to energy levels,
705: wave functions, and the PNC amplitude}
706: \label{sec:rad}
707:
708: The radiative corrections to the weak charge $Q_W$ have been
709: calculated for the free electron. However, an electron in a heavy
710: atom is bound, and this produces additional radiative
711: corrections proportional to $\alpha (Z \alpha)^n$,
712: $n=1,2,...$. Recently such corrections were considered
713: by Milstein and Sushkov \cite{milstein}. They found that
714: the most important are corrections enhanced by
715: the large parameter $\ln(\lambda/R)$, where $\lambda=\hbar/mc$
716: is the electron Compton wavelength and $R$ is the nuclear
717: radius. This type of correction arises from the radiative
718: corrections to the electron wave function near the nucleus.
719: In this region the $s$-wave and $p_{1/2}$-wave (lower Dirac component)
720: electron densities are singular, $|\psi(r)|^2 \sim r^{-Z^2\alpha^2}$.
721: The radiative corrections modify the potential at small distances
722: $r<\lambda$, ${\tilde V }(r)= -Z \alpha (1+\delta)/r$.
723: Correspondingly, the electron wave functions change,
724: $|\psi(r)|^2 \sim r^{-Z^2\alpha^2 (1 +\delta)^2}$ for $r<\lambda$.
725: This gives the radiative correction factor for the electron
726: density inside the nucleus,
727: \begin{equation}
728: \label{psirad}
729: \frac{|\psi(R)|^2}{|\psi(\lambda)|^2} \sim \Big( \frac{\lambda}{R}\Big)
730: ^{Z^2\alpha^2 2\delta}=\exp \Big( 2\delta Z^2\alpha^2 \ln(\lambda /R) \Big)
731: \ .
732: \end{equation}
733: For the Uehling (vacuum polarization) potential
734: $\delta \sim \alpha \ln(\lambda /r)$ \cite{berestetskii}.
735: This gives an additional power of the large parameter $\ln(\lambda / R)$.
736: This led Milstein and Sushkov \cite{milstein} to conclude that the
737: Uehling potential gives the dominating radiative correction
738: to $E_{PNC}$, $\sim Z^2\alpha^3 \ln^2(\lambda/ R)$.
739: Numerical calculations of the Uehling potential contribution
740: have been performed in \cite{johnson01} and in the present work.
741: This radiative correction increases $E_{PNC}$ by 0.4\%.
742:
743: Milstein and Sushkov \cite{milstein} demonstrated that there are no other
744: radiative corrections which are enhanced by $\ln^2(\lambda /R)$.
745: However, any correction to the potential with nonzero
746: $\delta(R) \sim \alpha$ gives a correction to the electron density
747: $\sim Z^2\alpha^3 \ln (\lambda /R)$. There are also corrections
748: $\sim Z^2\alpha^3 \ln (Z^2\alpha^2)$ which originate from the
749: shift of the energy levels (Lamb shift). We briefly discuss
750: these corrections below.
751:
752: Let us start our discussion from the radiative corrections to energy
753: levels.
754: The calculation of the shift can be divided into two parts:
755: one in which the electron interaction with virtual photons of
756: high-frequency are considered, and one in which
757: virtual photons of low-frequency are considered.
758:
759: In the high-frequency case the external field
760: (the strong nuclear Coulomb field) need only be included
761: to first-order.
762: In this case the contributions to the Lamb shift arise
763: from the diagrams presented in Fig.~\ref{fig:rad}.
764: The contribution of the Uehling potential (Fig.~\ref{fig:rad}(a))
765: to the Lamb shift is very small.
766: The main contribution comes from the vertex correction
767: (Fig.~\ref{fig:rad}(b)).
768: (In the case of a free electron the vertex diagrams give the electric
769: $f(q^{2})$ and magnetic $g(q^{2})$ form factors.)
770: The perturbation theory expression for $f(q^2)$ contains an
771: infra-red divergence and requires a low-frequency
772: cut-off parameter $\kappa$ - see, e.g., \cite{berestetskii}.
773: Assuming $q^2 \ll m^2 c^2$, the high-frequency contribution to the
774: Lamb shift can be presented as a potential given by the following
775: expression \cite{berestetskii}
776: \begin{eqnarray}
777: \delta \Phi ({\bf r})
778: &=&\Big[ \delta \Phi _{f}+\delta \Phi _{U}\Big] +\delta \Phi _{g}\nonumber \\
779: &=&\frac{\alpha \hbar ^{2}}{3\pi m^{2}c^{2}}\Big(
780: \ln \frac{m}{2\kappa} +\frac{11}{24}-\frac{1}{5}\Big)
781: \Delta \Phi ({\bf r})-
782: i\frac{\alpha \hbar}{4\pi mc}\mbox{\boldmath$\gamma$}\cdot
783: \mbox{\boldmath$\nabla$}\Phi ({\bf r}) \ .
784: \end{eqnarray}
785: For the Coulomb potential, $\Delta \Phi =-4\pi Ze\delta ({\bf r})$.
786: Here the last long-range term ($\delta \Phi _{g}$) comes from
787: the anomalous electron magnetic moment ($g(0)$);
788: the infra-red cut-off parameter $\kappa$ appears from the electric
789: form factor $f(q^{2})$.
790: This term with the large $\ln \frac{m}{2\kappa}$ gives the dominant
791: contribution to the Lamb shift of $s$-levels.
792: The infra-red divergence for $\kappa \rightarrow 0$ indicates the
793: importance of the low-frequency contribution for this term.
794:
795: If we go beyond the approximation
796: $q^{2}<<m^{2}c^{2}$, the term $\delta \Phi _{f}$ should be
797: associated with a non-local self-energy operator
798: ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',E)$ with typical values
799: $|{\bf r}-{\bf r}'|\lesssim \frac{\hbar}{mc}$
800: and $r\sim r' \lesssim \frac{\hbar}{mc}$.
801: We need this operator in a simple limit, $E<<mc^{2}$.
802: Matrix elements of ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',0)$
803: depend only on the electron density near the origin. Therefore,
804: we can express the Lamb shift of the external electron state in a
805: neutral atom in terms of the known Lamb shift
806: of highly excited states in hydrogen-like ions. To implement this
807: scheme we can approximate
808: ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',0)$
809: by a two-parametric $(A,b)$ potential
810: \begin{equation} \label{param}
811: \delta \Phi _{f}=-A\frac{\alpha}{\pi} r {\rm e}^{-b\frac{mc}{\hbar}r}\Phi \ ,
812: \end{equation}
813: where the factor $r$ is introduced to remove the singularity at $r=0$.
814: (We have also performed the calculation using the potential
815: $\delta \Phi _{f}$ without this factor $r$.
816: The results for the energy levels are the same.)
817: The parameters $A$ and $b$ in $\delta \Phi _{f}$ can be found from the
818: fit of the Lamb-shift of the high Coulomb levels $3s$, $4s$, $5s$ and
819: $3p$, $4p$ and $5p$ (in one-electron ions) which were calculated as
820: a function of the nuclear charge $Z$ in Refs. \cite{mohr}.
821: We have checked that $A_s=180$, $A_p=90$ and $b=1$ fit all these
822: Lamb shifts quite accurately
823: (for the potential without the factor $r$ we have found
824: $A_s=1.17$ and $A_p=1.33$).
825: The anomalous magnetic moment contribution is
826: \begin{equation}
827: \delta \Phi _{g}=-i\frac{\alpha \hbar}{4\pi mc}\mbox{\boldmath$\gamma$}
828: \cdot \mbox{\boldmath$\nabla$}\Phi \ .
829: \end{equation}
830: The Uehling potential for a finite nucleus is given by \cite{fullerton76}
831: (in atomic units: $\hbar =m=e=1$, $\alpha =1/c$)
832: \begin{equation}
833: \delta \Phi _{U}=-\frac{2\alpha ^{2}}{3r}\int _{0}^{\infty}dx~
834: x\rho(x)\int _{1}^{\infty}dt~\sqrt{t^{2}-1}
835: \Big(
836: \frac{1}{t^{3}}+\frac{1}{2t^{5}}\Big)
837: \Big(
838: {\rm e}^{-2t|r-x|/\alpha} -{\rm e}^{-2t(r+x)/\alpha} \Big) \ ,
839: \end{equation}
840: where $\rho (x)$ is the nuclear charge density.
841: It is more convenient to use a simpler formula for
842: $\delta \Phi _{U}$ for $r\geq R$,
843: \begin{eqnarray}
844: &&\delta \Phi _{U}(r)= \Phi (r)\frac{\alpha ^{4}}{8\pi R^{3}}
845: \int _{1}^{\infty}dt~ \sqrt{t^{2}-1}
846: \Big(
847: \frac{1}{t^{5}}+\frac{1}{2t^{7}}\Big)
848: {\rm e}^\frac{-2tr}{\alpha}I(x) \ , \\
849: &&I(x)=-{\rm e}^{x}+{\rm e}^{-x}+x{\rm e}^{x} +x{\rm e}^{-x} \ ,
850: \qquad x=2tR/\alpha \ ,
851: \end{eqnarray}
852: and take
853: $\frac{\delta \Phi _{U}(r<R)}{\Phi (r<R)}=
854: \frac{\delta \Phi _{U}(r=R)}{\Phi (r=R)}$.
855: There is practically no loss of numerical accuracy in this
856: approximation since a typical scale for the variation of
857: $\delta \Phi _{U}/\Phi$ is given by the electron Compton length
858: $\frac{\hbar}{mc}>>R$.
859:
860: Since $Z\alpha=0.4 $ is not so small for Cs it is important to estimate
861: the contribution of the higher-order in $Z\alpha$ vacuum-polarization
862: correction (the Wichmann-Kroll term \cite{kroll}). For simplicity we use
863: an approximate expression for this potential
864: \begin{equation}
865: \label{eq:WK}
866: \delta \Phi _{WK} = -\frac{2}{3}\frac{\alpha}{\pi} \,
867: \frac{0.092 Z^2 \alpha^2}{1+\big(\frac{1.62r}{\alpha}\big)^4}\Phi
868: \end{equation}
869: which is exact at small and large distances (a small-$r$ asymptotic was
870: presented in Refs. \cite{milstein83,milstein}).
871: We have found that the contribution of this potential to the $s$-wave
872: energies is $\sim 30$ times smaller than that of the Uehling potential.
873: The contribution of the Uehling potential $\delta \Phi _{U}$ to the Lamb
874: shift is always small (less than 15\%). Note that the contribution
875: of the radiative corrections to the electron core electrostatic
876: potential can be estimated using the semiclassical expression
877: for ${\hat \Sigma}_{\rm rad} ({\bf r},{\bf r}',E)$
878: presented in Ref. \cite{zelevinsky}. This contribution
879: for $s$-levels is two orders of magnitude smaller than
880: that of the nuclear potential
881: (for higher angular momenta the nuclear contribution is small and
882: the electron contribution is relatively important).
883:
884: The potential $\delta \Phi _{g}$ due to the magnetic form factor
885: is a long-range one.
886: This also hints that there are no large higher $Z\alpha$
887: corrections here. The contribution of this potential to the Lamb shift
888: is about 30\% for $s$-levels and 70\% for $p$-levels.
889: Note that for the term $\delta \Phi _{f}$ and the total Lamb-shift
890: we do not use the assumption $Z\alpha <<1$ since we fitted the exact results
891: for the single-electron ions.
892:
893: The radiative corrections for Cs energy levels are presented
894: in Table~\ref{tab:radenergies}.
895:
896: Now we discuss the contribution of
897: QED-type radiative corrections to the electron wave function
898: and PNC amplitude $E_{PNC}$.
899: Note that it is not enough to calculate the radiative corrections to
900: the matrix element of the weak interaction
901: $\langle n'p_{1/2}|\hat{H}_{W}|ns\rangle$.
902: Corrections to the energy intervals like $6s-6p$ are also important
903: since these intervals are small at the scale of the atomic unit
904: ($ \sim 1/20$) and sensitive to perturbations.
905: The change in the energies also influences the large-distance behavior
906: of the electron wave functions which determine the usual
907: E1 amplitudes in the sum-over-states approach.
908: If we keep only three dominating terms in the sum-over-states approach
909: (see below) the contributions of the energy shifts (-0.29 \%)
910: and corrections to the amplitudes (0.33\%) cancel each other.
911: The simplest way to find the total answer (including
912: the corrections to the weak matrix elements) is to include $\delta \Phi$
913: into the relativistic Hartree-Fock equations and
914: then perform all calculations.
915:
916: The results are the following:
917: the largest contribution to $E_{PNC}$ comes from the Uehling
918: potential $\delta \Phi _{U}$. It increases $E_{PNC}$ by
919: $0.41\%$ (in agreement with \cite{johnson01,milstein}).
920: The contribution of $\delta \Phi _{g}$ is $-0.03 \%$
921: (due to cancellation of the contributions of the corrections
922: to the $s$-wave and $p$-wave); the Wichmann-Kroll contribution is
923: $-0.006\%$. Note that the latter contribution can be estimated
924: analytically using Eq.~(\ref{psirad}). The ratio of the Wichmann-Kroll
925: contribution to the Uehling contribution in the logarithmic approximation is
926: $-0.184~Z^2\alpha^2/\ln (\lambda /R)$.
927:
928: The contribution of $\delta \Phi _{f}$ is well-defined only when
929: it originates from large distances (due to the shift of the energy levels
930: and the large-distance behavior of the electronic wave function
931: which influences electromagnetic amplitudes). A small-distance
932: contribution is not gauge-invariant and should be treated simultaneously
933: with the correction to the Z-boson exchange vertex
934: \cite{milstein}. Therefore, we deliberately selected a non-singular
935: parametric potential (\ref{param}) to approximate $\delta \Phi _{f}$.
936: This potential does not generate any singular
937: (in nuclear radius $R$) contributions proportional to the large
938: parameter $ \ln (\lambda /R)$. This means that the subject
939: of our discussion now is different from the radiative corrections
940: to the weak matrix element considered by Milstein and Sushkov
941: in \cite{milstein}. They were interested in contributions
942: enhanced by this large parameter $ \ln (\lambda /R)$.
943:
944: The main contribution to the Lamb shift comes from the distances
945: $r < 1/Z$ (in atomic units). This corresponds to the parameter
946: $b>Z\alpha $ in the potential (\ref{param}). On the other hand
947: a typical distance for the radiative corrections
948: is $r \sim \lambda=\frac{\hbar}{mc}$. Therefore, to estimate
949: the contribution of $\delta \Phi _{f}$ we performed calculations
950: for two extreme values of the parameters $b=1$ and $b=Z\alpha $
951: (the values of the Lamb shifts are kept the same in both cases
952: by selecting appropriate values of the parameter $A$).
953: For $b=1 $ this contribution is $-0.2\%$. For $b=Z\alpha $
954: it is -0.08\%. This gives us an indication that the large-distance
955: contribution to the radiative corrections to
956: $E_{PNC}$ is small. On the other hand, according to Sushkov and
957: Milstein \cite{milstein}, the small-distance
958: contribution to $E_{PNC}$ is dominated by the Uehling
959: potential which was considered above.
960:
961: %*****************************************************************
962: \section{Estimate of accuracy of PNC amplitude}
963: \label{sec:accuracy}
964:
965:
966: We have estimated the error of the PNC amplitude in a number of different
967: ways. There are two main methods:
968: (i) root-mean-square (rms) deviation of the calculated energy intervals,
969: E1 amplitudes, and hyperfine structure constants
970: from the accurate experimental values;
971: (ii) influence of fitting of energies and hyperfine structure
972: constants on the PNC amplitude.
973:
974: \subsection{Root-mean-square deviation}
975:
976: Remember that the PNC amplitude can be expressed as a sum over
977: intermediate states (see formula~(\ref{sum})).
978: Each term in the sum is a product of E1 transition amplitudes,
979: weak matrix elements, and energy denominators.
980: There are three dominating contributions to the $6s-7s$ PNC amplitude
981: in Cs:
982: \begin{eqnarray}
983: E_{PNC}&=&
984: \label{sumcs}
985: \frac{\langle 7s|\hat{H}_{E1}|6p\rangle \langle 6p|\hat{H}_{W}|6s\rangle}
986: {E_{6s}-E_{6p}} +
987: \frac{\langle 7s|\hat{H}_{W}|6p\rangle \langle 6p|\hat{H}_{E1}|6s\rangle}
988: {E_{7s}-E_{6p}}+
989: \frac{\langle 7s|\hat{H}_{E1}|7p\rangle \langle 7p|\hat{H}_{W}|6s\rangle}
990: {E_{6s}-E_{7p}}+...
991: \nonumber \\
992: &=& -1.908+1.493+1.352+...=0.937 +... \
993: \end{eqnarray}
994: (the numbers are from the work \cite{blundell90}
995: where the sum-over-states method was used;
996: here we just demonstrate that these terms dominate).
997: While we do not use the sum-over-states approach in our calculation of
998: the PNC amplitude, it is instructive to analyze the accuracy of
999: the E1 transition amplitudes, weak matrix elements, and
1000: energy intervals which contribute to Eq.~(\ref{sumcs}) as
1001: they have been calculated using the same method as that used to
1002: calculate $E_{PNC}$.
1003:
1004: Let us begin with the energy intervals.
1005: The calculated removal energies are presented in Table~\ref{tab:energies}.
1006: The Hartree-Fock values deviate from experiment by $10\%$.
1007: Including the second-order correlation corrections ${\hat \Sigma} ^{(2)}$
1008: reduces the error to $\sim 1\%$.
1009: When screening and the hole-particle interaction are included into
1010: ${\hat \Sigma} ^{(2)}$ in all orders,
1011: the energies improve, $\sim 0.3\%$.
1012: The percentage deviations from experiment of the energy intervals of interest
1013: are: $E_{6s}-E_{6p}$, $0.3$; $E_{7s}-E_{6p}$, $0.4$;
1014: and $E_{6s}-E_{7p}$, $0.3$. The rms error is $0.3\%$.
1015: We can in fact reproduce energy intervals exactly by placing coefficients
1016: before the correlation potential. We will use this procedure as another
1017: test of the stability of the results. Note, however, that the accuracy for
1018: the energies is already very high and the remaining discrepancy with
1019: experiment is of the same order of magnitude as the Breit and radiative
1020: corrections. Therefore, generally speaking, we should not expect that
1021: fitting of the energy will always improve the results for amplitudes and
1022: hyperfine structure. In fact, as we will see below, some values do improve
1023: while others do not. The overall accuracy, however, remains at the same level.
1024:
1025: The relevant E1 transition amplitudes (radial integrals) are presented in
1026: Table~\ref{tab:e1i}.
1027: These were calculated with the energy-fitted ``bare'' correlation
1028: potential $\hat{\Sigma}^{(2)}$ and the (unfitted and fitted)
1029: ``dressed'' potential $\hat{\Sigma}$.
1030: Structural radiation and normalization contributions were also
1031: included.
1032: In Table~\ref{tab:e1ii} the percentage deviations of
1033: the calculated values from experiment are listed.
1034: Without energy fitting, the rms error is $0.1\%$.
1035: Fitting the energy gives a rms error of $0.2\%$ for $\hat{\Sigma}^{(2)}$
1036: and $0.3\%$ for the complete $\hat{\Sigma}$.
1037:
1038: We cannot directly compare weak matrix elements with experiment.
1039: Like the weak matrix elements, hyperfine structure is determined
1040: by the electron wave functions in the vicinity of the nucleus,
1041: and this is known very accurately.
1042: The hyperfine structure constants calculated in different approximations
1043: are presented in Table~\ref{tab:hfsi}.
1044: Corrections due to the Breit interaction, structural radiation, and
1045: normalization are included.
1046: The percentage deviations from experiment are shown in
1047: Table~\ref{tab:hfsii}.
1048: The rms deviation of the calculated hfs values from experiment
1049: using the unfitted ${\hat \Sigma}$ is $0.5\%$.
1050: With fitting, the rms error in the pure second-order approximation is $0.3\%$;
1051: with higher orders we get $0.4\%$.
1052: We are, however, trying to estimate the accuracy of the $s-p$ weak
1053: matrix elements. It seems reasonable for us to use the square-root
1054: formula, $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$.
1055: The errors are presented in Table~\ref{tab:hfsii}.
1056: Notice that by using this approach the error is smaller.
1057: Without energy fitting, the rms error is $0.5\%$.
1058: With fitting, the rms error in the second-order calculation
1059: ($\hat{\Sigma} ^{(2)}$) is $0.2\%$ and in the full calculation
1060: ($\hat{\Sigma}$) it is $0.3\%$.
1061:
1062: From this section we can conclude that the rms error for the relevant
1063: parameters is $0.5\%$ or better.
1064:
1065: Note that from this analysis the error for the sum-over-states
1066: calculation of $E_{PNC}$ would be larger than this, as
1067: the errors for the energies, hfs constants, and E1 amplitudes
1068: contribute to each of the three terms in Eq.~(\ref{sumcs}).
1069: However, in the mixed-states approach, the errors do not add in this
1070: way.
1071: We get a better indication of the error of our calculation
1072: of $E_{PNC}$ in the next section.
1073:
1074:
1075: \subsection{Influence of fitting on the PNC amplitude}
1076:
1077: In the section above we presented calculations in three different
1078: approximations:
1079: with unfitted $\hat{\Sigma}$,
1080: and with $\hat{\Sigma}^{(2)}$ and $\hat{\Sigma}$ fitted with
1081: coefficients to reproduce experimental removal energies.
1082: The errors in these approximations are of different magnitudes and signs.
1083: We now calculate the PNC amplitude using these three approximations.
1084: The spread of the results can be used to estimate the error.
1085:
1086: The results are listed in Table~\ref{tab:pncii}.
1087: It can be seen that the PNC amplitude is very stable.
1088: The PNC amplitude is much more stable than hyperfine structure.
1089: This can be explained by the much smaller correlation corrections
1090: to $E_{PNC}$ ($\sim 2\%$ for $E_{PNC}$ and $\sim 30\%$ for hfs;
1091: compare Table~\ref{tab:pnci} with Table~\ref{tab:hfsi}).
1092: One can say that this small value of the correlation correction is
1093: a result of cancellation of different terms in (\ref{eq:pnc-cor})
1094: but each term is not small (see Table~\ref{tab:pnci}). However,
1095: this cancellation has a regular behavior. The same correlation
1096: potential $\hat \Sigma$ is used to calculate energies and correlation
1097: corrections (\ref{eq:pnc-cor}) to $E_{PNC}$. Therefore, whatever
1098: way is used to calculate $\hat \Sigma$, if the accuracy for energies
1099: is good, the correlation correction (\ref{eq:pnc-cor}) is stable.
1100: The stability of $E_{PNC}$ may be compared to the stability of the
1101: usual electromagnetic amplitudes where the error is very small
1102: (even without fitting).
1103:
1104: We have also considered the fitting of hyperfine structure using
1105: different coefficients before each $\hat{\Sigma}$.
1106: The first-order in $\hat{\Sigma}$ correlation correction (\ref{eq:pnc-cor})
1107: changes by about $10\%$. This changes the PNC amplitude
1108: by about $0.4\%$.
1109:
1110: \vspace{5mm}
1111:
1112: It is also instructive to look at the spread of $E_{PNC}$ obtained
1113: in different schemes.
1114: (This has already been discussed in some detail in
1115: Section~\ref{sec:results}.)
1116: The result of the present work is in excellent agreement with our earlier
1117: result \cite{dzuba89} while the calculation scheme is significantly different.
1118: The only other calculation of the $E_{PNC}$ in Cs which is as
1119: complete as ours is that of Blundell {\it et al.} \cite{blundell90}.
1120: Their result in the all-orders sum-over-states approach is 0.909
1121: (without Breit) and is very close to our value of 0.908
1122: (corresponding to ``Subtotal'' of Table~\ref{tab:pnci}).
1123: Our result (0.904) obtained in second-order with fitting of the energies
1124: is useful in determining the accuracy of the calculations of $E_{PNC}$.
1125: (Remember that this value is in agreement with results of similar
1126: calculations performed in \cite{blundell90,kozlov01}; see
1127: Section~\ref{sec:results}.) One can see that replacing
1128: the all-order $\hat \Sigma$ by its very rough second-order (with fitting)
1129: approximation changes $E_{PNC}$ by less than 0.4\% only. On the other hand,
1130: if the higher orders are included accurately, the difference between the
1131: two very different approaches is 0.1\% only.
1132:
1133: The maximum deviation we have obtained in the above analysis is
1134: $0.5\%$. We will therefore use this as our estimate for the
1135: uncertainty of the $E_{PNC}$ calculation.
1136:
1137: We do not include the error associated with the radiative corrections
1138: into the estimate of the accuracy since the large-distance
1139: ($r>\lambda$) contribution of the radiative corrections is small
1140: and the small-distance contribution ($R\lesssim r<\lambda$)
1141: enhanced by $\ln (\lambda/R)$ can be accurately calculated.
1142:
1143: %*****************************************************************
1144: \section{Conclusion}
1145:
1146: We have obtained the result
1147: \begin{equation}
1148: E_{PNC}=0.904\Big(1\pm 0.5\%\Big) \times 10^{-11}iea_{B}(-Q_{W}/N)
1149: \end{equation}
1150: for our calculation of the $6s-7s$ PNC amplitude in Cs.
1151: This is in agreement with other PNC calculations,
1152: however we would like to emphasize that our calculation is
1153: the most complete.
1154: The most precise measurement of the $6s-7s$ PNC amplitude in Cs
1155: is \cite{wood97}
1156: \begin{equation}
1157: -\frac{{\rm Im} (E_{PNC})}{\beta}=1.5939(56)\frac{\rm mV}{\rm cm} \ ,
1158: \end{equation}
1159: where $\beta$ is the vector transition polarizability.
1160: There are currently two very precise values for $\beta$. One value
1161: \begin{equation}
1162: \label{eq:beta1}
1163: \beta =26.957(51) a_{B}^{3}
1164: \end{equation}
1165: was obtained in our analysis \cite{dzuba00} of the Bennett and Wieman
1166: measurements \cite{wieman99} of the
1167: $M1_{\rm hfs}/\beta$ ratio \cite{bouchiat88}.
1168: We have obtained another value
1169: \begin{equation}
1170: \label{eq:beta2}
1171: \beta =27.15(11) a_{B}^{3}
1172: \end{equation}
1173: from the measurement \cite{cho97} of $\alpha/\beta$ and
1174: an analysis, similar to that of Ref. \cite{dzuba97}, using
1175: the most accurate experimental data for the E1 transition amplitudes
1176: including recent measurements of Vasilyev {\it et al.} \cite{vasilyev01}.
1177: The errors in Eqs.~(\ref{eq:beta1},\ref{eq:beta2}) are obtained by adding
1178: in quadrature the experimental and theoretical errors.
1179: Notice that the central point of our value (Eq.~(\ref{eq:beta2}))
1180: differs slightly from the value $27.22(11)$ obtained in the work
1181: \cite{vasilyev01}.
1182: Note also that the value (\ref{eq:beta2}) coincides
1183: with the value presented in \cite{dzuba97} which was obtained with slightly
1184: different E1 amplitudes. This is because of an accidental cancellation of
1185: the changes to different terms.
1186:
1187: Using the conversion $|e|/a_{B}^{2}=5.1422\times 10^{12}{\rm mV}/{\rm cm}$,
1188: we therefore obtain for the weak charge of the Cs nucleus
1189: with $\beta =26.957$:
1190: \begin{equation}
1191: \label{eq:q1}
1192: Q_{W}=-72.09(29)_{\rm exp}(36)_{\rm theor} \ ,
1193: \end{equation}
1194: or with $\beta =27.15$
1195: \begin{equation}
1196: \label{eq:q2}
1197: Q_{W}=-72.60(39)_{\rm exp}(36)_{\rm theor} \ ,
1198: \end{equation}
1199: where the experimental error is obtained by adding in quadrature the
1200: error for $\beta$ and the error for ${\rm Im}(E_{PNC})/\beta$.
1201:
1202: If we take an average value for $\beta$
1203: \begin{equation}
1204: \label{eq:beta3}
1205: \beta =26.99(5) a_{B}^{3}
1206: \end{equation}
1207: then
1208: \begin{equation}
1209: \label{eq:q3}
1210: Q_{W}=-72.18(29)_{\rm exp}(36)_{\rm theor} \ .
1211: \end{equation}
1212:
1213: These results (Eqs.~(\ref{eq:q1},\ref{eq:q2},\ref{eq:q3})) deviate by
1214: $2.2\sigma$, $0.9\sigma$ and $2.0\sigma$, respectively, from
1215: the standard model value $Q_{W}= -73.09(3)$ \cite{groom00}
1216: (note that the standard deviations $\sigma$ are different
1217: in each case).
1218:
1219:
1220: \acknowledgments
1221:
1222: We are grateful to A. Milstein, O. Sushkov, and M. Kuchiev for
1223: useful discussions.
1224: This work was supported by the Australian Research Council.
1225:
1226: %************************************************************
1227: \begin{thebibliography}{20}
1228:
1229: \bibitem{khriplovich}
1230:
1231: I.B. Khriplovich, {\it Parity Nonconservation in Atomic Phenomena}
1232: (Gordon and Breach, Philadelphia, 1991).
1233:
1234: \bibitem{bouchiat}
1235:
1236: M.-A. Bouchiat and C. Bouchiat, Rep. Prog. Phys. {\bf 60},
1237: 1351 (1997).
1238:
1239: \bibitem{wood97}
1240:
1241: C.S. Wood {\it et al.}, Science {\bf 275}, 1759 (1997).
1242:
1243: \bibitem{dzuba89}
1244:
1245: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
1246: Phys. Lett. A {\bf 141}, 147 (1989).
1247:
1248: \bibitem{blundell90}
1249:
1250: S.A. Blundell, W.R. Johnson, and J. Sapirstein,
1251: Phys. Rev. Lett. {\bf 65}, 1411 (1990);
1252: S.A. Blundell, J. Sapirstein, and W.R. Johnson,
1253: Phys. Rev. D {\bf 45}, 1602 (1992).
1254:
1255: \bibitem{rafac}
1256:
1257: R.J. Rafac, and C.E. Tanner, Phys. Rev. A, {\bf 58} 1087 (1998);
1258: R.J. Rafac, C.E. Tanner, A.E. Livingston, and H.G. Berry,
1259: Phys. Rev. A, {\bf 60} 3648 (1999).
1260:
1261: \bibitem{bennett99}
1262:
1263: S.C. Bennett, J.L. Roberts, and C.E. Wieman, Phys. Rev. A {\bf 59},
1264: R16 (1999).
1265:
1266: \bibitem{wieman99}
1267:
1268: S.C. Bennett and C.E. Wieman, Phys. Rev. Lett. {\bf 82},
1269: 2484 (1999); {\bf 82}, 4153(E) (1999); {\bf 83}, 889(E) (1999).
1270:
1271: \bibitem{RC}
1272:
1273: R. Casalbuoni, S. De Curtis, D. Dominici, and R. Gatto,
1274: Phys. Lett. B {\bf 460}, 135 (1999).
1275:
1276: \bibitem{JR}
1277:
1278: J. L. Rosner, Phys. Rev. D {\bf 61}, 016006 (1999).
1279:
1280: \bibitem{EL}
1281:
1282: J. Erler and P. Langacker, Phys. Rev. Letts. {\bf 84}, 212 (2000).
1283:
1284: \bibitem{derevianko00}
1285:
1286: A. Derevianko, Phys. Rev. Lett. {\bf 85}, 1618 (2000).
1287:
1288: \bibitem{harabati}
1289:
1290: V.A. Dzuba, C. Harabati, W.R. Johnson, and M.S. Safronova,
1291: Phys. Rev. A {\bf 63}, 044103 (2001).
1292:
1293: \bibitem{kozlov01}
1294:
1295: M.G. Kozlov, S.G. Porsev, and I.I. Tupitsyn,
1296: Phys. Rev. Lett. {\bf 86}, 3260 (2001).
1297:
1298: \bibitem{marciano}
1299:
1300: W.J. Marciano and A. Sirlin, Phys. Rev. D {\bf 27}, 552 (1983);
1301: W.J. Marciano and J.L. Rosner, Phys. Rev. Lett. {\bf 65}, 2963 (1990).
1302:
1303: \bibitem{lynnbed}
1304:
1305: B.W. Lynn and P.G.H. Sandars, J. Phys. B. {\bf 27}, 1469 (1994);
1306: I. Bednyakov {\it et al.}, Phys. Rev. A {\bf 61}, 012103 (1999).
1307:
1308: \bibitem{milstein}
1309:
1310: A.I. Milstein and O.P. Sushkov, e-print hep-ph/0109257.
1311:
1312: \bibitem{johnson01}
1313:
1314: W.R. Johnson, I. Bednyakov, and G. Soff,
1315: Phys. Rev. Lett. {\bf 87}, 233001 (2001).
1316:
1317: \bibitem{derevianko}
1318:
1319: A. Derevianko, Phys. Rev. A {\bf 65}, 012106 (2002).
1320:
1321: \bibitem{dzuba00}
1322:
1323: V.A. Dzuba and V.V. Flambaum,
1324: Phys. Rev. A {\bf 62}, 052101 (2000).
1325:
1326: \bibitem{dzuba97}
1327:
1328: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
1329: Phys. Rev. A {\bf 56}, R4357 (1997).
1330:
1331: \bibitem{groom00}
1332:
1333: D.E. Groom {\it et al.},
1334: Euro. Phys. J. C {\bf 15}, 1 (2000).
1335:
1336: \bibitem{dzuba87}
1337:
1338: V.A. Dzuba, V.V. Flambaum, P.G. Silvestrov, and O.P. Sushkov,
1339: J. Phys. B {\bf 20}, 1399 (1987).
1340:
1341: \bibitem{dzuba89energy}
1342:
1343: V.A. Dzuba, V.V. Flambaum, and O.P. Sushkov,
1344: Phys. Lett. A {\bf 140}, 493 (1989).
1345:
1346: \bibitem{dzuba89e1hfs}
1347:
1348: V.A. Dzuba, V.V. Flambaum, A.Ya. Kraftmakher, and O.P. Sushkov,
1349: Phys. Lett. A {\bf 142}, 373 (1989).
1350:
1351: %\bibitem{dzuba}
1352: %
1353: %V.A. Dzuba, in preparation.
1354:
1355: %-------------
1356: %proton distrn
1357:
1358: \bibitem{fricke95}
1359:
1360: G. Fricke {\it et al.},
1361: At. Data and Nucl. Data Tables {\bf 60}, 177 (1995).
1362:
1363: %-------------
1364: %neutron distrn
1365:
1366: \bibitem{r_{np}}
1367:
1368: A. Trzci\'{n}ska {\it et al.},
1369: Phys. Rev. Lett. {\bf 87}, 082501 (2001).
1370:
1371: \bibitem{berestetskii}
1372:
1373: V.B. Berestetskii, E.M. Lifshitz, and L.P. Pitaevskii,
1374: {\it Relativistic Quantum Theory}
1375: (Pergamon Press, Oxford, 1982).
1376:
1377: \bibitem{mohr}
1378:
1379: P.J. Mohr and Y.-K. Kim, Phys. Rev. A {\bf 45}, 2727 (1992);
1380: P.J. Mohr, Phys. Rev. A {\bf 46}, 4421 (1992).
1381:
1382: \bibitem{fullerton76}
1383:
1384: L.W. Fullerton and G.A. Rinker, Jr., Phys. Rev. A {\bf 13},
1385: 1283 (1976).
1386:
1387: \bibitem{kroll}
1388:
1389: E.H. Wichmann and N.M. Kroll,
1390: Phys. Rev. {\bf 101}, 343 (1956).
1391:
1392: \bibitem{milstein83}
1393:
1394: A.I. Milstein and V.M. Strakhovenko,
1395: ZhETF {\bf 84}, 1247 (1983).
1396:
1397: \bibitem{zelevinsky}
1398:
1399: V.V. Flambaum and V.G. Zelevinsky,
1400: Phys. Rev. Lett. {\bf 83}, 3108 (1999).
1401:
1402: %------------------
1403: % energies
1404:
1405: \bibitem{moore}
1406:
1407: C.E. Moore, Natl. Stand. Ref. Data Ser.
1408: (U.S., Natl. Bur. Stand.), {\bf 3} (1971).
1409:
1410: %------------------
1411: % radial integrals
1412:
1413: \bibitem{rafac99}
1414:
1415: R.J. Rafac, C.E. Tanner, A.E. Livingston, and H.G. Berry,
1416: Phys. Rev. A {\bf 60}, 3648 (1999).
1417:
1418: \bibitem{bouchiat84}
1419:
1420: M.-A. Bouchiat, J. Gu\'{e}na, and L. Pottier,
1421: J. Phys. (France) Lett. {\bf 45}, L523 (1984).
1422:
1423: %------------------
1424: % hfs
1425:
1426: \bibitem{cshfs}
1427:
1428: E. Arimondo, M. Inguscio, and P. Violino,
1429: Rev. Mod. Phys. {\bf 49}, 31 (1977).
1430:
1431: \bibitem{gilbert83}
1432:
1433: S.L. Gilbert, R.N. Watts, and C.E. Wieman,
1434: Phys. Rev. A {\bf 27}, 581 (1983).
1435:
1436: \bibitem{rafac97}
1437:
1438: R.J. Rafac and C.E. Tanner, Phys. Rev. A {\bf 56}, 1027 (1997).
1439:
1440: %---------------------
1441:
1442: \bibitem{bouchiat88}
1443:
1444: M.-A. Bouchiat and J. Gu\'{e}na, J. Phys. (France) {\bf 49},
1445: 2037 (1988).
1446:
1447: \bibitem{cho97}
1448:
1449: D. Cho {\it et al.}, Phys. Rev. A {\bf 55}, 1007 (1997).
1450:
1451: \bibitem{vasilyev01}
1452:
1453: A.A. Vasilyev, I.M. Savukov, M.S. Safronova, and H.G. Berry,
1454: e-print physics/0112071.
1455:
1456:
1457: \end{thebibliography}
1458:
1459: %************************************************************
1460: \begin{table}
1461: \caption{Contributions to the $6s-7s$ $E_{PNC}$ amplitude
1462: for Cs in units $10^{-11}iea_{B}(-Q_{W}/N)$.
1463: ($\hat{\Sigma}$ corresponds to the (unfitted) ``dressed''
1464: self-energy operator.) }
1465: \label{tab:pnci}
1466: \begin{tabular}{ld}
1467: TDHF & 0.8898 \\
1468: $\langle \psi _{7s}|\hat \Sigma_s(\epsilon_{7s})|\delta X_{6s}\rangle$
1469: & 0.0773 \\
1470: $\langle \delta\psi _{7s}|\hat \Sigma_p(\epsilon_{7s})|X_{6s}\rangle$
1471: & 0.1799 \\
1472: $\langle \delta Y_{7s}|\hat \Sigma_s(\epsilon_{6s})|\psi_{6s}\rangle$
1473: & -0.0810 \\
1474: $\langle Y_{7s}|\hat \Sigma_p(\epsilon_{6s})|\delta \psi_{6s}\rangle$
1475: & -0.1369 \\
1476: Nonlinear in $\hat{\Sigma}$ correction & -0.0214 \\
1477: Weak correlation potential & 0.0038 \\
1478: Structural radiation & 0.0029 \\
1479: Normalization & -0.0066 \\
1480: & \\
1481: Subtotal & 0.9078 \\
1482: & \\
1483: Breit & -0.0055 \\
1484: Neutron distribution & -0.0018 \\
1485: Radiative corrections & 0.0036 \\
1486: & \\
1487: Total & 0.9041 \\
1488: \end{tabular}
1489: \end{table}
1490: %****************************************************************
1491: \begin{table}
1492: \caption{Radiative corrections to RHF removal energies; units cm$^{-1}$.
1493: (See also Table~\ref{tab:energies}.) }
1494: \label{tab:radenergies}
1495: \begin{tabular}{dddd}
1496: $6s$ & $7s$ & $6p_{1/2}$ & $7p_{1/2}$ \\
1497: \hline
1498:
1499: -18.4 & -5.0 & 0.88 & 0.31 \\
1500: \end{tabular}
1501: \end{table}
1502: %**********************************************************
1503: \begin{table}
1504: \caption{Removal energies for Cs in units cm$^{-1}$.}
1505: \label{tab:energies}
1506: \begin{tabular}{lllll}
1507: State & RHF & $\hat{\Sigma} ^{(2)}$ & $\hat{\Sigma}$ &
1508: Experiment \tablenotemark[1]\\
1509: \hline
1510: $6s$ & 27954 & 32415 & 31492 & 31407 \\
1511: $7s$ & 12112 & 13070 & 12893 & 12871 \\
1512: $6p_{1/2}$ & 18790 & 20539 & 20280 & 20228 \\
1513: $7p_{1/2}$ & 9223 & 9731 & 9663 & 9641 \\
1514: \end{tabular}
1515: \tablenotetext[1]{Taken from \cite{moore}.}
1516: \end{table}
1517: %************************************************************
1518: \begin{table}
1519: \caption{Radial integrals of E1 transition amplitudes for Cs in
1520: different approximations. The experimental values are listed in
1521: the last column. (a.u.)}
1522: \label{tab:e1i}
1523: \begin{tabular}{ldddddd}
1524: Transition & RHF& TDHF & $\hat{\Sigma} ^{(2)}$ &
1525: $\hat{\Sigma}$ & $\hat{\Sigma}$ & Experiment \\
1526: & & & with fitting & & with fitting &
1527: \\
1528: \hline
1529: $6s-6p$ & 6.464 & 6.093 & 5.499 & 5.497 & 5.509 &
1530: 5.497(8) \tablenotemark[1] \\
1531: $7s-6p$ & 5.405 & 5.450 & 5.198 & 5.190 & 5.204 &
1532: 5.185(27) \tablenotemark[2]\\
1533: $7s-7p$ & 13.483 & 13.376 & 12.602 & 12.601 & 12.612 &
1534: 12.625(18) \tablenotemark[3] \\
1535: \end{tabular}
1536: \tablenotetext[1]{Ref. \cite{rafac99}.}
1537: \tablenotetext[2]{Ref. \cite{bouchiat84}.}
1538: \tablenotetext[3]{Ref. \cite{bennett99}.}
1539: \end{table}
1540: %************************************************************
1541: \begin{table}
1542: \caption{Percentage deviation from experiment of calculated
1543: E1 transition amplitudes in different approximations.}
1544: \label{tab:e1ii}
1545: \begin{tabular}{lddd}
1546: Transition &
1547: \multicolumn{3}{c}{Percentage deviation} \\
1548: & $\hat{\Sigma}^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1549: & with fitting & & with fitting \\
1550: \hline
1551: $6s-6p$ & 0.04 & 0.0 & 0.2 \\
1552: $7s-6p$ & 0.3 & 0.1 & 0.4 \\
1553: $7s-7p$ & -0.2 & -0.2 & -0.1 \\
1554: \end{tabular}
1555: \end{table}
1556: %************************************************************
1557: \begin{table}
1558: \caption{Calculations of the hyperfine structure of Cs
1559: in different approximations. In the last column the experimental
1560: values are listed. Units: MHz.}
1561: \label{tab:hfsi}
1562: \begin{tabular}{ldddddd}
1563: State & RHF & TDHF & $\hat{\Sigma} ^{(2)}$ &
1564: $\hat{\Sigma}$ & $\hat{\Sigma}$ & Experiment \\
1565: & & & with fitting & & with fitting & \\
1566: \hline
1567: $6s$ & 1425.0 & 1717.5 & 2306.9 & 2315.0 & 2300.3 &
1568: 2298.2 \tablenotemark[1] \\
1569: $7s$ & 391.6 & 471.1 & 544.4 & 545.3 & 543.8 &
1570: 545.90(9) \tablenotemark[2] \\
1571: $6p_{1/2}$ & 160.9 & 200.3 & 291.5 & 293.6 & 290.5 &
1572: 291.89(8) \tablenotemark[3] \\
1573: $7p_{1/2}$ & 57.6 & 71.2 & 94.3 & 94.8 & 94.1 &
1574: 94.35 \tablenotemark[1] \\
1575: \end{tabular}
1576: \tablenotetext[1]{Ref. \cite{cshfs}.}
1577: \tablenotetext[2]{Ref. \cite{gilbert83}.}
1578: \tablenotetext[3]{Ref. \cite{rafac97}.}
1579: \end{table}
1580: %************************************************************
1581: \begin{table}
1582: \caption{Percentage deviation from experiment of calculated
1583: hyperfine structure constants in different approximations.}
1584: \label{tab:hfsii}
1585: \begin{tabular}{ldddd}
1586: State &
1587: \multicolumn{3}{c}{Percentage deviation} \\
1588: & $\hat{\Sigma}^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1589: & with fitting & & with fitting \\
1590: \hline
1591: $6s$ & 0.4 & 0.7 & 0.09 \\
1592: $7s$ & -0.3 & -0.1 & -0.4 \\
1593: $6p_{1/2}$ & -0.1 & 0.6 & -0.5 \\
1594: $7p_{1/2}$ & -0.05 & 0.5 & -0.3 \\
1595: \end{tabular}
1596: \end{table}
1597: %****************************************************************
1598: \begin{table}
1599: \caption{Percentage deviation from experiment of calculated
1600: $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$ (we will denote this by $s-p$ in
1601: the tables) in different approximations.}
1602: \label{tab:hfsiii}
1603: \begin{tabular}{lddd}
1604: $\sqrt{{\rm hfs}(s){\rm hfs}(p)}$&\multicolumn{3}{c}{Percentage deviation}\\
1605: & $\hat{\Sigma} ^{(2)}$ & $\hat{\Sigma}$ & $\hat{\Sigma}$ \\
1606: & with fitting & & with fitting \\
1607: \hline
1608: $6s-6p$ & 0.1 & 0.7 & -0.2 \\
1609: $6s-7p$ & 0.2 & 0.6 & -0.09 \\
1610: $7s-6p$ & -0.2 & 0.2 & -0.4 \\
1611: \end{tabular}
1612: \end{table}
1613: %****************************************************************
1614: \begin{table}
1615: \caption{Values for $E_{PNC}$ in different approximations;
1616: units $10^{-11}iea_{B}(-Q_{W}/N)$.}
1617: \label{tab:pncii}
1618: \begin{tabular}{cddd}
1619: & $\hat{\Sigma}^{(2)}$ with fitting & $\hat{\Sigma}$ &
1620: $\hat{\Sigma}$ with fitting \\
1621: \hline
1622: $E_{PNC}$ & 0.901 & 0.904 & 0.903 \\
1623: \end{tabular}
1624: \end{table}
1625: %****************************************************************
1626:
1627: \center
1628: \widetext
1629: \input psfig
1630: \psfull
1631:
1632: \begin{figure}[b]
1633: \centerline{\psfig{file=Sigma2.eps,clip=}}
1634: \caption{Second-order (a) direct and (b) exchange correlation diagrams
1635: in the Feynman diagram technique.
1636: Dashed line is the Coulomb interaction.
1637: Loop is the polarization of the atomic core.}
1638: \label{fig:Sigma2}
1639: \end{figure}
1640:
1641: \begin{figure}[b]
1642: \centerline{\psfig{file=hpchain.eps, clip=}}
1643: \caption{Hole-particle interaction in the polarization operator.}
1644: \label{fig:h-p}
1645: \end{figure}
1646:
1647: \begin{figure}[b]
1648: \centerline{\psfig{file=Sigmad.eps,clip=}}
1649: \caption{Screening of the Coulomb interaction
1650: in the direct correlation diagram.}
1651: \label{fig:Sigmad}
1652: \end{figure}
1653:
1654: \begin{figure}[b]
1655: \centerline{\psfig{file=screeninghp.eps,clip=}}
1656: \caption{The screened Coulomb interaction
1657: (with hole-particle interaction).}
1658: \label{fig:screeninghp}
1659: \end{figure}
1660:
1661: \begin{figure}[b]
1662: \centerline{\psfig{file=Sigmahpsc.eps, clip=}}
1663: \caption{The electron self-energy operator with screening and hole-particle
1664: interaction included.}
1665: \label{fig:Sigmahpsc}
1666: \end{figure}
1667:
1668: \begin{figure}[b]
1669: \centerline{\psfig{file=pncdom.eps, clip=}}
1670: \caption{Lowest-order correlation corrections to the PNC E1 transition
1671: amplitude. Dashed line is the E1 field; cross is the nuclear weak field.}
1672: \label{fig:pncdom}
1673: \end{figure}
1674:
1675: \begin{figure}[b]
1676: \centerline{\psfig{file=pncint.eps, clip=}}
1677: \caption{Small corrections to the PNC E1 transition amplitude:
1678: external field inside the correlation potential.
1679: In diagrams (a) the weak interaction is inside the
1680: correlation potential ($\delta \hat{\Sigma}$ denotes the change in
1681: $\hat{\Sigma}$ due to the weak interaction);
1682: this is known as the weak correlation potential.
1683: Diagrams (b) represent structural radiation
1684: (photon field inside the correlation potential).}
1685: \label{fig:pncint}
1686: \end{figure}
1687:
1688: \begin{figure}[b]
1689: \centerline{\psfig{file=rad2.eps, clip=}}
1690: \caption{High-frequency contribution to radiative corrections.
1691: Diagram (a) corresponds to the Uehling potential.
1692: Diagram (b) is the vertex correction.
1693: The single solid line represents the bound electron;
1694: the double line is the free electron;
1695: the Coulomb interaction is denoted by the dashed line;
1696: and the filled circle denotes the nucleus.}
1697: \label{fig:rad}
1698: \end{figure}
1699:
1700: \end{document}
1701:
1702:
1703: